A Short Essay on Variational Calculus

Similar documents
Robotics. Islam S. M. Khalil. November 15, German University in Cairo

HAMILTON S PRINCIPLE

Sketchy Notes on Lagrangian and Hamiltonian Mechanics

Introduction to the Calculus of Variations

Optimization using Calculus. Optimization of Functions of Multiple Variables subject to Equality Constraints

Numerical Optimization Algorithms

Physics 5153 Classical Mechanics. Canonical Transformations-1

PH.D. PRELIMINARY EXAMINATION MATHEMATICS

General Variation of a Functional

Physics 129a Calculus of Variations Frank Porter Revision

The Principle of Least Action

Lecture 10: Solving the Time-Independent Schrödinger Equation. 1 Stationary States 1. 2 Solving for Energy Eigenstates 3

Page 684. Lecture 40: Coordinate Transformations: Time Transformations Date Revised: 2009/02/02 Date Given: 2009/02/02

Advanced Classical Mechanics (NS-350B) - Lecture 3. Alessandro Grelli Institute for Subatomic Physics

Under evolution for a small time δt the area A(t) = q p evolves into an area

PH.D. PRELIMINARY EXAMINATION MATHEMATICS

for changing independent variables. Most simply for a function f(x) the Legendre transformation f(x) B(s) takes the form B(s) = xs f(x) with s = df

Lecture 9: Eigenvalues and Eigenvectors in Classical Mechanics (See Section 3.12 in Boas)

AN INTRODUCTION TO LAGRANGE EQUATIONS. Professor J. Kim Vandiver October 28, 2016

Legendre Transforms, Calculus of Varations, and Mechanics Principles

Analytical Dynamics: Lagrange s Equation and its Application A Brief Introduction

Physics 200 Lecture 4. Integration. Lecture 4. Physics 200 Laboratory

Section 9 Variational Method. Page 492

Johann Bernoulli ( )he established the principle of the so-called virtual work for static systems.

Lecture 4. Alexey Boyarsky. October 6, 2015

Classical mechanics of particles and fields

Analytical Dynamics: Lagrange s Equation and its Application A Brief Introduction

Curves in the configuration space Q or in the velocity phase space Ω satisfying the Euler-Lagrange (EL) equations,

The Particle-Field Hamiltonian

Strauss PDEs 2e: Section Exercise 1 Page 1 of 6

Calculus of Variations

Lecture 20: Modes in a crystal and continuum

Functional differentiation

Coordinate systems and vectors in three spatial dimensions

5.1 Classical Harmonic Oscillator

df(x) = h(x) dx Chemistry 4531 Mathematical Preliminaries Spring 2009 I. A Primer on Differential Equations Order of differential equation

We start with some important background material in classical and quantum mechanics.

Calculus of Variations Summer Term 2014

An Investigation of the Attainable Efficiency of Flight at Mach One or Just Beyond

Massachusetts Institute of Technology Department of Physics. Final Examination December 17, 2004

Calculus of Variations

Physics 6010, Fall 2016 Constraints and Lagrange Multipliers. Relevant Sections in Text:

Harmonic Oscillator I

Mathematical Methods of Physics I ChaosBook.org/ predrag/courses/phys Homework 1

Forces of Constraint & Lagrange Multipliers

RELATING CLASSICAL AND QUANTUM MECHANICS 1 I: Thinking about the wave function

Virtual Work and Variational Principles

Physical meaning of Lagrange multipliers

COMPLETE ALL ROUGH WORKINGS IN THE ANSWER BOOK AND CROSS THROUGH ANY WORK WHICH IS NOT TO BE ASSESSED.

Quantum Field Theory III

harmonic oscillator in quantum mechanics

An introduction to Variational calculus in Machine Learning

Sec. 1.1: Basics of Vectors

REVIEW. Hamilton s principle. based on FW-18. Variational statement of mechanics: (for conservative forces) action Equivalent to Newton s laws!

Hamiltonian. March 30, 2013

Principles of Optimal Control Spring 2008

arxiv:physics/ v1 [physics.ed-ph] 18 Oct 2004

Variational Methods and Optimal Control

Chapter 1. Variational Method

Gauge Fixing and Constrained Dynamics in Numerical Relativity

Second Order Optimality Conditions for Constrained Nonlinear Programming

STABILITY. Phase portraits and local stability

Calculus of Variations Summer Term 2014

Problem set 2: Solutions Math 207B, Winter 2016

Mechanics IV: Oscillations

VIII. Phase Transformations. Lecture 38: Nucleation and Spinodal Decomposition

Gaussian integrals and Feynman diagrams. February 28

5 Handling Constraints

Deriving Lagrange s Equations Using Elementary Calculus

7 EQUATIONS OF MOTION FOR AN INVISCID FLUID

Attempts at relativistic QM

The Path Integral Formulation of Quantum Mechanics

Brachistochrones with loose ends

1 Quantum fields in Minkowski spacetime

Optimization. Escuela de Ingeniería Informática de Oviedo. (Dpto. de Matemáticas-UniOvi) Numerical Computation Optimization 1 / 30

GENERAL RELATIVITY: THE FIELD THEORY APPROACH

The first order formalism and the transition to the

(Most of the material presented in this chapter is taken from Thornton and Marion, Chap. 6)

Chapter 1. Principles of Motion in Invariantive Mechanics

Numerical Methods of Applied Mathematics -- II Spring 2009

Classical Mechanics in Hamiltonian Form

Problem 1: Lagrangians and Conserved Quantities. Consider the following action for a particle of mass m moving in one dimension

How to Simulate a Trebuchet Part 1: Lagrange s Equations

Non-Linear Response of Test Mass to External Forces and Arbitrary Motion of Suspension Point

Lecture 19: Calculus of Variations II - Lagrangian

General elastic beam with an elastic foundation

Large Deviations for Small-Noise Stochastic Differential Equations

Introduction to Vector Functions

L = 1 2 a(q) q2 V (q).

EE291E/ME 290Q Lecture Notes 8. Optimal Control and Dynamic Games

Noether s Theorem. 4.1 Ignorable Coordinates

Optimization Tutorial 1. Basic Gradient Descent

Resonant Tunneling in Quantum Field Theory

Lagrange Multipliers Method

Physics 141 Energy 1 Page 1. Energy 1

Chapter 2 The Euler Lagrange Equations and Noether s Theorem

Numerical Methods for Partial Differential Equations

etc., etc. Consequently, the Euler Lagrange equations for the Φ and Φ fields may be written in a manifestly covariant form as L Φ = m 2 Φ, (S.

Quantum Field Theory II

1 Mathematical preliminaries

Transcription:

A Short Essay on Variational Calculus Keonwook Kang, Chris Weinberger and Wei Cai Department of Mechanical Engineering, Stanford University Stanford, CA 94305-4040 May 3, 2006 Contents 1 Definition of a functional 1 2 First variation 2 3 Essential and natural boundary conditions 4 4 Second variation 5 5 Functional derivative 6 6 Examples 7 6.1 Lagrange s equation of classical mechanics........................ 7 6.2 Tight string under lateral loading............................. 8 6.3 Phase field equation for microstructure evolution.................... 9 6.4 Schrödinger s equation of quantum mechanics...................... 9 1 Definition of a functional The calculus of variations deals with functionals, which are functions of a function, to put it simply. For example, the calculus of variations can be used to find an unknown function that minimizes or maximizes a functional. Let us begin our discussion with the definition of a functional. We are familiar with the definition of a function. It is a mapping from one number (or a set of numbers) to another value. For example, f(x) = x 2 (1) is a function, which maps x = 2 to f(x) = 4, and x = 3 to f(x) = 9, etc. On the other hand, a functional is a mapping from a function (or a set of functions) to a value. For example, π 0 [y(x)] 2 dx is a functional. Notice that we use square brackets for a functional, to signify the fact that its argument is a function. When y(x) = x, π 0 x 2 dx = π3 3 1 (2) (3)

When y(x) = sin x, π 0 sin 2 x dx = π 2 Therefore, functional I[y(x)] maps y(x) = x to π 3 /3 and maps y(x) = sin x to π/2. Because an integral maps a function to a number, a functional usually involves an integral. The following form of functional often appears in the calculus of variations, F (x, y(x), y (x)) dx (5) where y (x) dy(x)/dx. Many engineering problems can be formulated into a variational form, in which we need to find a function y(x) that minimizes (or maximizes) this functional, subject to certain boundary conditions, such as y( ) = a and y(x 2 ) = b. Notice that while I[y(x)] is a functional, F (x, y, y ) is simply a function that maps three numbers into one number. 2 First variation For a function f(x), its differential, df, is how much f changes if its argument, x, changes by an infinitesimal amount dx. For example, when f(x) = x 2, df = 2x dx If x is at a minimum (or maximum) of f(x), then df should be zero for an infinitesimal change of x. In this example, the minimum occurs at x = 0. For a functional I[y(x)], the corresponding term is its (first) variation, δi. δi is how much I changes if its argument, the function y(x), changes by an infinitesimal amount δy(x). This is illustrated in Fig. 1, where function y(x) is shown in bold line. If the function undergoes a small change, ỹ(x) = y(x) + δy(x) (7) where δy(x) is a small, continuous function, then the variation of the functional is, δi = I[ỹ(x)] I[y(x)] (8) Our task is obtain the explicit expression of δi. Suppose that the functional I[y(x)] takes the form of Eq.(5). To obtain δi, let us define δy(x) = ɛφ(x) (9) where φ(x) is an arbitrary continuous function whose value is of the order 1 and ɛ is an infinitesimal number. We also require φ(x) to be sufficiently smooth so that ỹ (x) is well defined. 1 We can now Taylor expand I[ỹ(x)] in terms of ɛ. I[ỹ(x)] = = x 1 x2 = I[y(x)] + ɛ F (x, y(x) + ɛφ(x), y (x) + ɛφ (x)) dx [ F (x, y(x), y (x)) + [ ɛφ(x) + φ(x) + φ (x) ] ɛφ (x) + O(ɛ 2 ) ] dx (4) (6) dx + O(ɛ 2 ) (10) 1 Another restriction on φ(x) is that φ(x) and φ (x) have the same order of smallness. In this case, the variation in y is said to be weak [2]. 2

y δy(x) ~ y(x) y(x) B A x 2 x Figure 1: Function y(x) (bold line) and a small variation from it, ỹ(x) = y(x) + δy(x). δy(x) is a small, continuous function which satisfies the boundary conditions δy( ) = δy(x 2 ) = 0. Keeping terms only up to the first order, we have, [ ] δi = ɛ φ(x) + φ (x) dx [ ] = δy(x) + δy (x) dx (11) Usually, it is more desirable to express δi solely in terms of δy(x), and get rid of δy (x) in the expression. This can be done by integration-by-parts. δi = ɛ = ɛ x 1 x2 φ(x) dx + ɛ dφ(x) [ ] x2 φ(x) dx + ɛ φ(x) ɛ ( ) d dx φ(x) dx (12) If we constrain the value of function y(x) at and x 2 to be constant, i.e., apply boundary conditions φ( ) = φ(x 2 ) = 0, then, 2 [ δi = ɛ d ( )] dx φ(x) dx [ = d ( )] dx δy(x) dx (13) Therefore, if y(x) is a minimum (or a maximum) of functional I[y(x)], subjected to the constraint that y( ) and y(x 2 ) are fixed, then y(x) must satisfy the differential equation, d ( ) dx = 0 (14) for x (, x 2 ). This is called the Euler-Lagrange equation. 2 The same conclusion can be made if y(x) satisfies periodic boundary conditions, y() = y(x 2), y () = y (x 2). 3

Example. What is the shortest path between two points A and B, as shown in in Fig.1? Of course, the answer is a straight line. Here we will derive this result from variational calculus. Given a curve y(x) connecting points A and B, the length of the curve is, L[y(x)] = F (x, y(x), y (x)) dx = 1 + (y (x)) 2 dx (15) Therefore, L is a functional of y(x). The minimum of functional L must satisfy the Euler-Lagrange equation (14). Notice that in this case F has no explicit dependence on x or y(x), so that, = 0 and = The Euler-Lagrange equation becomes ( ) d y (x) = 0 d + (y (x)) 2 y (x) 1 + (y (x)) 2 The solution is that y (x)=const, hence the shortest path is a straight line. 3 Essential and natural boundary conditions In the above derivation of the Euler-Lagrange equation, we assumed that δy( ) = δy(x 2 ) = 0. This makes the second term in Eq. (12) vanish. In this case, the function y(x) satisfies the boundary condition, y( ) = y 1 y(x 2 ) = y 2 (16) (17) where y 1 and y 2 are constants. This is called the essential (or geometric) boundary condition. This is the appropriate boundary condition for some applications, such as the example discussed above. However, in other applications, we may need to apply other types of boundary conditions to the function y(x). If we still want the second term in Eq. (12) to vanish (so that we obtain the familiar Euler- Lagrange equation), but allowing δy( ) and δy(x 2 ) to be non-zero, then we need to have, = 0 (18) x=x1 = 0 (19) x=x2 This is called a natural boundary condition. A system may also have a natural boundary condition at one end (x = ) and an essential boundary condition at the other end (x = x 2 ). If the functional involves higher derivatives of y(x), such as, F (x, y(x), y (x), y (x)) dx (20) 4

then the minimum (or maximum) of I[y(x)] satisfies the following Euler-Lagrange equation, d 2 ( ) dx 2 d ( ) dx + = 0 (21) provided that y(x) is subjected to the following essential or natural boundary conditions, essential y( ) = y 1 y ( ) = y 1 y(x 2 ) = y 2 y (x 2 ) = y 2 natural d dx ( ) = 0 at x = = 0 at x = d dx ( ) = 0 at x = x 2 = 0 at x = x 2 In general, when the functional I contains higher derivatives of y(x), i=0 the Euler-Lagrange equation is n ( ) ( 1) i di dx i (i) = 0 F (x, y, y, y,, y (n) ) dx (22) when appropriate boundary condition is applied. 4 Second variation When y(x) satisfies the Euler-Lagrange equation, the first variation δi vanishes and y(x) is an extremum of the functional I[y(x)]. To find out whether y(x) is a minimum or a maximum of I[y(x)], we need to look at the second variation of I. Let us continue the Taylor expansion in Eq. (10) to the second order, I[ỹ(x)] = = I[y(x)] + ɛ + ɛ2 2 F (x, y(x) + ɛφ(x), y (x) + ɛφ (x)) dx [ ] φ(x) + φ (x) dx [ 2 F 2 φ2 (x) + 2 F 2 φ 2 (x) + 2 F φ(x)φ (x) ] dx + O(ɛ 3 ) (23) The second variation of I is the term that is of the second order of ɛ, δ 2 I = ɛ2 x2 [ 2 ] F 2 2 φ2 (x) + 2 F 2 φ 2 (x) + 2 F φ(x)φ (x) dx = 1 [ 2 ] F 2 2 δy2 (x) + 2 F 2 δy 2 (x) + 2 F δy(x)δy (x) dx (24) When the first variation δi vanishes, y(x) is a maximum if the second variation δ 2 I > 0 (for arbitrary φ(x)), or a minimum if δ 2 I < 0. If δ 2 I = 0, we will have to examine the higher order variations of I to determine whether y(x) is a minimum or a maximum. For more discussions see [2, 3]. 5

5 Functional derivative For a function with multiple arguments, f(, x 2,, x n ), if the differential df can be written as, df = n g i (,, x n ) dx i (25) i=1 then g i (,, x n ) is called the (partial) derivative of f with respect to x i, for i = 1,, n, f x i = g i (,, x n ) (26) Notice that the derivative of a function f(,, x n ) with respect to x i g i (,, x n ). Similarly, if the (first) variation of a functional is another function b a can be written as, F (y(x), y (x)) dx (27) δi = b a g(x) δy(x) dx then the functional derivative of I is (28) δi = g(x) (29) δy(x) The functional derivative of a functional I[y(x)] is a function g(x). In the following, we discuss the connection between the (ordinary) derivative and the functional derivative. Imagine that we approximate function y(x) by a piece-wise linear curve that passes through a set of points (x i, y i ), i = 0,, n + 1, where x i = i, y i = y(x i ), x 0 = a, and x n+1 = b, as shown in Fig. 2. Let y(x) subject to essential boundary conditions, so that y 0 = y(a) and y n+1 = y(b) are fixed. Then the functional I[y(x)] can be approximated by a sum, where b a = lim n I n (y 1, y 2,, y n ) F (y(x), y (x)) dx n i=0 ( F y i, y ) i+1 y i lim n I n(y 1,, y n ) (30) n F (y i, z i ) (31) i=0 z i y i+1 y i (32) 6

y δy k δy k a... y k+1 x 0 x 2 x 3 x k-1 x k x k+1 x n+1 y k... b x Figure 2: A piece-wise linear approximation of function y(x). x i = i, y i = y(x i ), x 0 = a, x n+1 = b. Consider that one of the node, (x k, y k ) moves vertically by δy k, the variational derivative δi/δy at x = x k is the change of I divided by the change of area δy (shade region) in the limit of 0. Therefore, functional I[y(x)] is the limit of an ordinary function I n (y 1,, y n ) when the number of its arguments, n, goes to infinity. The partial derivative of I n with respect to one of its arguments is, I n k = ( [ + 1 y=yk d y=yk dx 1 y =z k 1 ( ) ] x=x k y =z k Compare this with the definition of the functional derivative, we have, δi δy(x) = d ( ) I n 1 dx = lim 0 k ) (33) Therefore, the functional derivative δi/δy(x) can be interpreted as the following. Imagine a small and localized perturbation of y(x) near point x, δi/δy(x) is the change of I divided by the change of area under the curve y(x). 6 Examples 6.1 Lagrange s equation of classical mechanics The equations of motion for particles in classical mechanics can be obtained from Hamilton s principle or the principle of least action. Let q be the coordinate of a particle, t be the time, so that q(t) is the trajectory of this particle. Suppose that the particle s position at times t 1 and t 2 are known, (34) q(t 1 ) = q 1 q(t 2 ) = q 2 (35) (36) i.e., q(t) is subjected to essential boundary conditions. The question is: what trajectory q(t) would the particle take to go through points q 1 and q 2 exactly at times t 1 and t 2? 7

The Hamilton s principle states that the real trajectory of the particle is the one that minimizes (or maximizes) the action functional [5], S[q(t)] = t2 t 1 L(t, q(t), q (t)) dt (37) where L(t, q(t), q (t)) is the Lagrangian of the system, which is usually the kinetic energy minus the potential energy, L(q, q ) = 1 2 m(q (t)) 2 V (q) (38) The real trajectory of the particle must have δs = 0, and hence satisfies the Euler-Lagrange equation, ( ) d L dt q L q = 0. (39) When the Lagrangian takes the form of Eq. (38), this leads to the following equation of motion, mq (t) = V q (40) which is identical to the Newton s equation of motion F = ma, where F = V q is the force, a = q (t) is the acceleration. In general, if the system has n degrees of freedom, q 1,, q n, the Lagrange s equations of motion are ( ) d L L = 0 for all i = 1,2,,n. dt q i q i 6.2 Tight string under lateral loading When a system subjected to external loading reaches thermal equilibrium, its Gibbs free energy is minimized. This thermodynamic principle allows us to determine the equilibrium shape of an elastic body subjected to external loading. For example, consider an elastic string subjected to lateral loading w(x), as shown in Fig. 3. Ignoring the entropic contribution, the Gibbs free energy of the system is simply the enthalpy H, i.e. the store elastic energy minus the work done by the external load. Assuming the string has a line tension T, the enthalpy can be written as the following functional of the shape of the string y(x), (in the limit of y (x) 1) H[y(x)] = L 0 Therefore, the variational derivative is 1 2 T (y (x)) 2 w(x) y(x) dx (41) δh δy(x) = T y (x) w(x) (42) Because the equilibrium shape of the string minimizes the functional H[y(x)], it satisfies the following Euler-Lagrange equation, T y (x) + w(x) = 0 (43) 8

0 w(x) y(x) L x y Figure 3: Distributed load w(x) is applied to the string hanging tightly between two walls. The deflection y(x) and its derivative is assumed to be very small because the string is tight. 6.3 Phase field equation for microstructure evolution Many systems evolves in the direction that minimizes its Gibbs free energy. This observation can be used to construct evolution equations of a wide range of systems, such as the phase field model of material microstructure [7]. For example, consider a simple model of crystal growth by solidification from the melt. The system may be described by a phase field, φ(x), such that φ(x) 1 if the material at x is solid and φ(x) 1 if it is liquid. The free energy F of the system is a functional of the phase field φ(x), which can be written as, F [φ(x)] = L 0 U(φ(x) 1) 2 (φ(x) + 1) 2 + ɛ (φ (x)) 2 dx (44) The first term in the integrand accounts for the system s preference in either one of the two states, φ = ±1. The second term in the integrand introduces a free energy penalty for the liquid-solid interface. A simple evolution equation for φ(x) is that φ(x) d dtφ(x) follows the steepest descent direction, i.e. the direction of the variational derivative, φ(x) = k δf δφ(x) = k [ 4Uφ(x) (φ(x) 2 1) ɛ φ (x) ] (45) where k is a rate constant. 6.4 Schrödinger s equation of quantum mechanics An electron in quantum mechanics is described by its wave function ψ(x). For a single electron moving in a potential V (x), the Hamiltonian operator is, 2 Ĥ = 2 + V (x) (46) 2m x2 This means that, Ĥψ(x) = 2 2 ψ(x) 2m x 2 + V (x)ψ(x) (47) The energy E of the electron is a functional of its wave function ψ(x), E[ψ(x)] = ψ (x)ĥψ(x) dx (48) 9

For simplicity, we will assume that ψ(x) is real, so that its complex conjugate ψ (x) = ψ(x). We will also assume that ψ( ) = ψ( ) = 0, i.e. assume essential boundary conditions. The ground state wave function is the minimizer of functional E[ψ(x)]. However, the wave function must satisfy the normalization condition, ψ (x)ψ(x) dx = 1 (49) Therefore, the ground state wave function satisfies the following Euler-Lagrange equation, where δe δψ(x) = 0 (50) E [ψ(x)] = E[ψ(x)] λ ψ (x)ψ(x) dx (51) and λ is the Lagrange multiplier. The explicit expression of the variational derivative is, δe [ δψ(x) = 2 2 2 ] ψ(x) 2m x 2 + V (x) ψ(x) λ ψ(x) (52) Therefore, the ground state wave function satisfies the following eigen equation, Ĥψ(x) = λψ(x) (53) We can further show that λ = E[ψ(x)] in this case, so that the constant λ is usually written as E. This leads to the time-independent Schrödinger s equation, [ 2 2 ] 2m x 2 + V (x) ψ(x) = E ψ(x) (54) References [1] R. S. Schechter, The Variational Method in Engineering, (McGrow-Hill Book Company, 1967). [2] C. Fox, An Introduction to the Calculus of Variations, (Oxford University Press, 1950). [3] I. M. Gelfand and S. V. Fomin, Calculus of Variations, (Dover, New York, 2000). [4] B. van Brunt, The Calculus of Variations, (Springer, 2004). [5] L. D. Landau and E. M. Lifshitz, Mechanics, (Pergamon Press, 1969). [6] C. Lanczos, The Variational Principles of Mechanics, (Dover, New York, 1970).h 9 [7] V. V. Bulatov and W. Cai, Computer Simulation of Dislocations, (Oxford University Press, to be published, 2006), chapter 11. 10